Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

drones

Article
Drone-Action: An Outdoor Recorded Drone Video
Dataset for Action Recognition
Asanka G. Perera 1, * , Yee Wei Law 1 and Javaan Chahl 1,2
1 School of Engineering, University of South Australia, Adelaide SA 5095, Australia;
yeewei.law@unisa.edu.au (Y.W.L.); Javaan.Chahl@unisa.edu.au (J.C.)
2 Joint and Operations Analysis Division, Defence Science and Technology Group,
Melbourne, VIC 3207, Australia
* Correspondence: asanka.perera@mymail.unisa.edu.au

Received: 6 November 2019; Accepted: 24 November 2019; Published: 28 November 2019 

Abstract: Aerial human action recognition is an emerging topic in drone applications. Commercial
drone platforms capable of detecting basic human actions such as hand gestures have been developed.
However, a limited number of aerial video datasets are available to support increased research into
aerial human action analysis. Most of the datasets are confined to indoor scenes or object tracking and
many outdoor datasets do not have sufficient human body details to apply state-of-the-art machine
learning techniques. To fill this gap and enable research in wider application areas, we present an
action recognition dataset recorded in an outdoor setting. A free flying drone was used to record 13
dynamic human actions. The dataset contains 240 high-definition video clips consisting of 66,919
frames. All of the videos were recorded from low-altitude and at low speed to capture the maximum
human pose details with relatively high resolution. This dataset should be useful to many research
areas, including action recognition, surveillance, situational awareness, and gait analysis. To test
the dataset, we evaluated the dataset with a pose-based convolutional neural network (P-CNN)
and high-level pose feature (HLPF) descriptors. The overall baseline action recognition accuracy
calculated using P-CNN was 75.92%.

Keywords: drone; dataset; human action recognition; aerial video analysis; P-CNN

1. Introduction
Drones or Unmanned aerial vehicles (UAVs) are increasingly popular due to their affordability
and applicability in numerous commercial applications. Some popular application areas of drones
are photogrammetry [1], agriculture [2], crowd monitoring [3], sports activity recording [3], parcel
delivery [3], and search and rescue [4,5]. Recently, these areas have experienced rapid advances
through convergence of technologies [6].
For research in these areas, a series of datasets have been released in the past few years. These
datasets cover multiple research disciplines but mainly in the security, industrial, and agricultural
sectors. Examples of such application-specific drone datasets include datasets for object detection [7,8],
datasets for vehicle trajectory estimation [9,10], datasets for object tracking [11,12], datasets for human
action recognition [13–16], datasets for gesture recognition [17–19], datasets for face recognition [20,21],
a dataset for fault detection in photovoltaic plants [22], datasets for geographic information system [23,24],
and datasets for agriculture [25,26].
However, limited research has been done on human–drone interactions and associated
applications. Although some commercial drones can recognize gestures, the capabilities rely on
high-quality videos captured at low altitudes and are limited to a small number of simple gestures.
Going beyond gestures, enabling a drone to recognize general human actions is far more challenging.

Drones 2019, 3, 82; doi:10.3390/drones3040082 www.mdpi.com/journal/drones


Drones 2019, 3, 82 2 of 16

A limited number of studies have been conducted to understand the sophisticated human body
movements recorded from a drone [16]. Reasons for the slow progress in human action recognition in
aerial videos include:

• Human action recognition is inherently a challenging problem. Most of the action recognition
studies are focused on popular video datasets that consist mostly of ground-level videos [27].
Even when using high quality videos, the state of the art in action recognition is still fallible [28].
• The quality of aerial videos is often marred by a lack of image details, perspective distortion,
occlusions, and camera movements.
• Many drone applications require online rather than offline processing, but the resource constraints
of embedded hardware platforms limit the choice and complexity of action recognition algorithms.
• There are not enough relevant video datasets to help train algorithms for action recognition in
the aerial domain. Currently available aerial video datasets are mostly limited to object tracking.
While there are some datasets supporting research in aerial action recognition [16], they are limited.

Based on the above, we reason that action recognition using aerial videos is less studied than
general human action recognition. Recognizing the importance of datasets for action recognition
research in the aerial domain, we present in this study, a new full high-definition (FHD) video dataset
with rich human details, that was recorded on a drone in a controlled manner.
The work presented here is specifically focused on providing accurate and rich human body
details in aerial videos recorded from a moving platform. State-of-the-art action recognition techniques
mostly rely on high-resolution human videos to achieve high accuracy, typically by leveraging by
computationally intensive techniques. However, many real-life aerial videos suffer from low resolution
and undesired camera movement resulting from the apparatus. We believe that current limitations
of aerial camera systems should not deprive action recognition research of high-quality and highly
detailed aerial videos.
The dataset presented here consists of 13 action classes recorded at FHD (1920 × 1080 resolution)
from a quadrotor drone flying slowly at low altitude. Some actions were recorded while the drone
was hovering (e.g., kicking, stabbing, and punching), and others were recorded while the drone was
following the subject (e.g., walking, jogging, and running). A hovering camera should not be confused
with a stationary camera. A hovering camera can drift in any direction due to platform mobility and
susceptibility to wind gusts. All of the videos were recorded in a way that preserved as much body
surface area as possible. This dataset was designed to support research related to search and rescue,
situational awareness, surveillance, pose estimation, and action recognition. We assume that in most
practical missions, a drone operator or an autonomous drone follows these general rules:

• Avoid flying at overly low altitude, which is hazardous to humans and equipment;
• Avoid flying at overly high altitude, so as to maintain sufficient image resolution;
• Avoid high-speed flying, and therefore, motion blur;
• Hover to acquire more details of interesting scenes;
• Record human subjects from a viewpoint that gives minimum perspective distortion.

Our dataset was created assuming the guidelines above were followed.
For testing the dataset, a pose-based convolutional neural network (P-CNN) [29] was used as
the baseline action recognition algorithm. P-CNN utilizes the CNN features of body parts extracted
using the estimated pose. The dataset was also evaluated with high-level pose features (HLPF) [30]
which recognize action classes based on the temporal relationships of body joints and their variations.
The baseline accuracy and the experiment details have been compared with recently published human
activity datasets.
The rest of this article is organized as follows. Section 2 discusses closely related work on popular
action recognition datasets, aerial video datasets, and their limitations in terms of suitability for
aerial human action recognition. Section 3 describes the steps involved in preparing the dataset,
Drones 2019, 3, 82 3 of 16

and compares the dataset with other recently published video datasets. Section 4 reports experimental
results. A discussion of issues and potential improvements is presented in Section 5. Section 6
concludes. Throughout this manuscript we will use the noun “drone” synonymously with the acronym
UAV which describes an aircraft in terms of not being manned.

2. Related Work
A number of video datasets have been published for general human action recognition and aerial
human action recognition. Here, we briefly discuss some studies related to our work.
Human action recognition datasets (mostly focused on ground videos): KTH [31] and
Weizmann [32] were the most popular early action recognition datasets, and they were recorded
in a controlled setting with a static background. These datasets helped to progress action recognition
research. UCF101 [33] was one of the largest and most diverse datasets in this category. It was created
with 13,320 YouTube videos belonging to 101 action classes. JHMDB [30] and Penn action [34] are
another two notable action datasets created using YouTube videos. These datasets come with human
pose and action annotations.
The recently made action recognition datasets are massive in terms of their numbers of videos and
action classes. Sports-1M [35] and ActivityNet [36] were two datasets introduced under this category.
These datasets do not provide spatial localization details of human subjects. For example, Sports-1M
videos are annotated automatically using YouTube topics and are, therefore, weakly labeled [37].
The most recent additions to this category were YouTube-8M [38] Kinetics [39], and HACS [40].
The Kinetics dataset consisted of 400, 600, and 700 action classes in its successive releases. HACS
dataset was created with 1.55 million annotated clips across 200 classes. The action classes of both
datasets cover a diverse range of daily human actions collected from YouTube videos. Diversified large
datasets are useful for pre-training action recognition networks on them and enhancing networks’
performances on small datasets.
Aerial datasets: The available aerial video datasets can be broadly categorized as follows based
on their intended applications.
• Object detection and tracking datasets: There has been a surge of interest in aerial object
detection and tracking studies. They are mainly focused on vehicle and human detection and
tracking. VisDrone [11] is the largest object detection and tracking dataset in this category. This
dataset covers various urban and suburban areas with a diverse range of aerial objects (it includes
2.5 million annotated instances). The dataset has been arranged into four tracks for object detection
in images/videos and single/multi object tracking. The similar but relatively smaller UAVDT
dataset [41] contains 80 thousand annotated instances recorded from low altitude drones. UAV123
dataset [15] has been presented for single object tracking from 123 aerial videos. Videos in these
datasets have been annotated for the ground truth object bounding boxes.
• Gesture recognition datasets: A gesture dataset recorded from a low-altitude hovering drone was
presented in [17]. The UAV-Gesture dataset contains outdoor recorded of 13 drone commanding
signals. NATOPS [42] is a similar indoor recorded gesture dataset consisting of 24 aircraft
signal gestures.
• Human action recognition datasets: A large-scale VIRAT dataset [13] contains 550 videos
covering a range of realistic and controlled human actions. The dataset has been recorded from
both static and moving cameras (called VIRAT ground and aerial datasets). There are 23 outdoor
event types with 29 h of videos. A limitation of the VIRAT aerial dataset is its low resolution of
480 × 720 pixels restricts algorithms from retrieving rich activity information from relatively small
human subjects. A 4k resolution Okutama-Action [16] video dataset was introduced to detect
12 concurrent actions by multiple subjects. The dataset was recorded in a baseball field using
two drones. Abrupt camera movements are present in the dataset making it more challenging
for action recognition. However, the 90 degree elevation angle of the camera creates severe
self-occlusions and perspective distortions in the videos. Other notable datasets in this category
Drones 2019, 3, 82 4 of 16

are UCFARG [14], UCF aerial action [43], and Mini-drone [44]. UCF aerial and UCF ARG datasets
have been recorded from an R/C-controlled blimp and a helium balloon respectively. Both
datasets have similar action classes. UCF ARG is a multi-view dataset with synchronized ground,
rooftop, and aerial camera views of the actions. However, UCF aerial action is a single-view
dataset. Mini-drone was developed to study various aspects and definitions of privacy in aerial
videos. The dataset was captured in a car park covering illicit, suspicious, and normal behaviors.

Most state-of-the-art human action recognition systems are implemented in high-end or graphics
processing unit (GPU) powered computers, and those algorithms cannot perform as intended on typical
embedded platforms due to the slowing of gains in commodity processor performance despite strong
market pull [45]. The platform dependency and the limitations of processing power in drones have
been bottlenecks to the development of onboard machine learning and AI techniques for human action
or gesture recognition. Notably, action recognition from a mobile platform is not an extensively studied
research area. Some notable works related to aerial image analysis are limited to offline processing
of recorded videos [9,46–49]. In order to utilize the drone’s mobile and operational capabilities to
meet the emerging demand for drone technology, more studies are needed. In particular, the aerial
action recognition studies need more datasets to develop technologies tailored to search and rescue,
surveillance, and situational awareness. We tried to collect the data against relatively clutter-free
backgrounds that were nonetheless normal outdoor environments.

3. Preparing the Dataset


This section explains the data collection process, the selected action classes, and the variations in
the collected data. The dataset is compared with recently published aerial and ground video datasets.

3.1. Data Collection


The data collection was done on an unsettled road located in the middle of a wheat field. The aerial
platform was a 3DR SOLO rotorcraft. The flight was slow and done at low altitudes (8–12 m). Videos
were recorded using a GoPro Hero 4 Black camera with an anti-fish eye replacement lens (5.4 mm,
10 MP, IR CUT) and a 3-axis Solo gimbal. All videos in the dataset are HD (1920 × 1080) format
captured at 25 fps.
A total of 13 actions were recorded while the drone was in hovering and following modes. In all
videos, the subject was roughly in the middle of the frame and performed each action five to ten times
except for walking, jogging, and running. For walking, jogging, and running actions, the subject was
asked to traverse between two markers while the drone followed the subject from different directions.
When recording the actions, sometimes the drone drifted from its initial hovering position due
to wind gusts. This adds random and characteristic camera motion to the videos corresponding to
practical scenarios.
The total number of human subjects in the dataset was 10. The participants were asked to perform
the actions in a selected section of the road. We divided the actions into the following three categories
based on the viewpoint and the camera movement: following, side-view, and front-view. A detailed
illustration of these three settings is provided in Figure 1.
Following actions: For the walking, jogging, and running actions, the subjects were recorded from
four viewpoints. Each action was repeated four times while recording from the drone positioned at
four different viewpoints in succession. As shown in Figure 1a, the left and right side views were
recorded by maintaining a roughly perpendicular view with respect to the walking direction. However,
due to mismatches in the moving speeds of the subject and the drone, in some videos, the subject was
not always in the middle of the frame. Irrespective of the horizontal and vertical movements of the
subject with respect to the camera frame, the videos were trimmed to include only the full body details
of the person undertaking the particular action. We tried to follow and record the subject throughout
the action, but in some videos, the drone was in hover mode or hover and follow modes combined.
Drones 2019, 3, 82 5 of 16

In particular, when the subject was running fast, the recordings were mostly done in the hovering
mode or had a little movement towards the direction in which the subject was moving. In terms of the
camera motion, the recorded actions are from both moving and hovering modes.
Side-view actions: Five actions were recorded from both left and right sides of the subject
when the drone was in hover mode; namely, punching, hitting with a bottle, hitting with a stick, kicking,
and stabbing. We used mock foam weapons to represent sticks, bottles, and knives. Each action was
repeated five to ten times (see Figure 1b). The actions were recorded with diversity across the class.
For example, for the kicking action, each subject performed kicks towards the left and right sides of the
camera frame. Each instance (left or right) had five to ten randomly performed kicks.
Front-view actions: Clapping and waving hands actions were recorded while the drone was
hovering in front of the subject (see Figure 1c). In these videos, the subject was roughly in the middle
of the frame and repeated each action five to ten times.

(a) (b) (c)

Figure 1. The actions were recorded in three different settings using a single drone. They are
(a) following actions, (b) side-view actions, and (c) front-view actions.

3.2. Action Class Selection


The action categories described in Section 3.1 were selected considering three application scenarios.
(i) Analyzing human movements: This includes the following actions (walking, jogging, and running).
Each action was recorded from four viewpoints from the front, back, and left and right sides of the
subject. In the dataset, we consider the action recorded from the front and back views as one class and
the action recorded from left and right sides as another class. For example, the walking_side action class
has 20 videos recorded from left and right sides of 10 subjects, while walking_front_back has 20 videos
recorded from front and back viewpoints. The 120 videos in these six action classes are also suitable
for gait analysis and tracking.
(ii) Violence detection: The five side-view actions were selected from common violent acts. Punching
and kicking are the most common bare hand actions of humans when fighting with each other [50].
It is also common to use some objects to hit or throw at each other [51]. To cover these scenarios, we
included hitting with a bottle, hitting with a stick, and stabbing. Each action was recorded from left and
right sides of the subject, capturing the maximum body surface area possible of these asymmetric
actions. There are 100 videos showing the violent actions of 10 subjects.
(iii) Signaling the drone: In some scenarios, people might try to signal to the drone for help.
The most popular way of signaling someone at a visible distance is using hand gestures. We selected
two actions (clapping and waving hands) for this scenario and recorded from the front view as front-view
actions (Figure 2). Each class contains 10 videos.
Drones 2019, 3, 82 6 of 16

Walking f/b Walking side Jogging f/b Jogging side Running f/b

Running side Hitting with bottle Hitting with stick Stabbing

Punching Kicking Clapping Waving hands

Figure 2. The action classes of the Drone-Action dataset. The images shown were selected randomly
and cropped around the subject for clear demonstration.

3.3. Variations in Data


The actors were asked to perform each action five to ten times, with most of the actions being
performed ten times. Each actor performed the actions differently, as they wished, capturing a variety
of natural responses. All actors were volunteers from our research group and consisted of seven males
and three females.
There are rich variations in the recorded actions in terms of the phase, orientation, camera
movement, and the body shapes of the actors. The camera movements were caused by wind gusts and
by the movement of the drone. In some videos, the scale rapidly changes. An example of scale change
is the running action recorded from front or back viewpoints. The drone could not always maintain
the speed of the subject in low-altitude flight. This resulted in scale variation of subjects during the
course of the video. In some videos, the skin color of the actor is close to the background color. These
variations create a challenging dataset for action recognition, and also makes it more representative of
real-world situations.

3.4. Dataset Annotations


All of the videos have been annotated with subject ID, action class, and bounding box. Subject
IDs are S1 to S10 for the ten actors. Along with bounding box annotations, body joint estimations
computed using the widely used pose estimator OpenPose [52] were also included.
The experiments conducted here used three randomly generated split sets. These split sets were
generated using the subject IDs associated with the videos; i.e., by avoiding overlapping of subject IDs
between the train and test sets. Subject IDs can be used to generate customized split sets. The split sets
used in our experiments are available for download with the dataset.
Drones 2019, 3, 82 7 of 16

3.5. Dataset Summary


The dataset contains a total of 240 clips with 66919 frames. The number of actors in the dataset is
10 and they performed each action 5–10 times. All the videos are provided with 1920×1080 resolution
and 25 fps. The average duration of each action was 11.15 sec. A summary of the dataset is given
in Table 1. The total clip length and mean clip length of each class are represented on the left side
(blue) and right side (amber) bar graphs of Figure 3 respectively. In Table 2, we compare our dataset
with eight recently published video datasets. These datasets have helped progress research in action
recognition, gesture recognition, event recognition, and object tracking.

Table 1. A summary of the dataset.

Feature Value
# Actions 13
# Actors 10
# Clips 240
# Clips per class 10-20
Repetitions per class 5-10
Mean clip length 11.15 sec
Total duration 44.6 mins
# Frames 66919
Frame rate 25 fps
Resolution 1920 × 1080
Camera motion Yes (hover and follow)
Annotation Bounding box

Table 2. Comparison of recently published video datasets.

Dataset Scenario Purpose Environment Frames Classes Resolution Year


UT Surveillance Action Outdoor 36k 6 360 × 240 2010
Interaction [53] recognition
NATOPS [42] Aircraft Gesture Indoor N/A 24 320 × 240 2011
signaling recognition
VIRAT [13] Drone, Event Outdoor Many 23 Varying 2011
surveillance recognition
UCF101 [33] YouTube Action Varying 558k 24 320 × 240 2012
recognition
J-HMDB [30] Movies, Action Varying 32k 21 320 × 240 2013
YouTube recognition
Mini-drone [44] Drone Privacy Outdoor 23.3k 3 1920 × 1080 2015
protection
Campus [54] Surveillance Object Outdoor 11.2k 1 1414 × 2019 2016
tracking
Okutama- Drone Action Outdoor 70k 13 3840 × 2160 2017
Action [16] recognition
UAV-Gesture [17] Drone Gesture Outdoor 37.2k 13 1920 × 1080 2018
recognition
Drone-Action Drone Action Outdoor 66.9k 13 1920 × 1080 2019
recognition
Drones 2019, 3, 82 8 of 16

350 20

Mean clip length (sec)


300

Total clip length (sec)


250 15

200
10
150
100 5
50
0 0

Pu ing

W king /b
Ki e

St ide
g

al ng
g

g de

s
in /b
gg ick

ni f/b
w le

nd
d
in

un in

f
tt

f
si

W bi

i
t

ck

s
pp

R ch

s
W king
itt . bo

Jo ing

un g
.s

ha
ab
R nin
g

ng
n
la
w
C

gg

in
g
Jo

al
g
in

av
in
itt
H
H

Figure 3. The total clip length (blue) and the mean clip length (amber) are shown in the same graph
in seconds.

4. Experimental Results
We experimented with two popular feature types used in human action recognition; namely,
pose-based Cheron et al.’s CNN features [29] and Jhuang et al.’s high-level pose features [30]. In this
section, we report and compare experimental results using these two approaches.

4.1. High-Level Pose Features (HLPF)


HLPFs are formed by combining spatial and temporal properties of body keypoints throughout the
action. We used the publicly available HLPF code [30] with minor modifications. HLPF was calculated
using 15 keypoints (head, neck, shoulders, elbows, wrists, abdomen, hips, knees, and ankles). We
used the pose estimator OpenPose [52] to find keypoints (see Figure 4a,b). In each frame, four spatial
properties were calculated as follows:

• Normalized positions: Each key point was normalized with respect to the belly key point. A total of
30 descriptors were obtained (x and y coordinates of 15 joints).
• Distance relationships: The distance between two key points was calculated for all 15 key points,
resulting in 105 descriptors.
• Angular relationships: A total of 1365 angles were calculated. Each angle was that of two vectors
connecting a triplet of key points.
• Orientation relationships: 104 angles were calculated between each vector connecting two key
points and the vector from neck to abdomen.

The temporal features were obtained by combining the following trajectories:

• Cartesian trajectory: 30 descriptors were obtained from the translation of 15 key points along x and
y axes.
• Radial trajectory: The radial displacement of 15 key points provides 15 descriptors.
• Distance relationship trajectory: 105 descriptors were calculated from the differences between
distance relationships.
• Angle relationship trajectory: 1365 descriptors were calculated from the differences between
angle relationships.
• Orientation relationship trajectory: 104 descriptors were calculated from the differences between
distance relationships.

An HLPF vector (including 3223 descriptor types) was generated by combining the above spatial
and temporal descriptors. A codebook was generated by performing k-means clustering on each
Drones 2019, 3, 82 9 of 16

descriptor type. The codeword size was selected to have k = 20. Classification was done using a SVM
with linear kernels.

(a) (b) (c)

Left hand Right hand Upper body Full body Full image

RGB
patches

Flow
patches

(d)

Figure 4. Extracted part patches (left hand, right hand, upper body, and full body) are shown using
an image from the kicking class. (a) The estimated pose is shown with the background. The image
is cropped around the subject for clear demonstration. (b) The keypoint locations are used compute
high-level pose features (HLPF) and posed-based convolutional neural network (cP-CNN) features.
(c) Part locations were determined based-on the keypoints. Purple color bounding box covers the
full image. (d) RGB and flow patches were extracted from the RGB images and optical flow images
respectively, using the bounding boxes shown in (c).

4.2. Pose-Based CNN (P-CNN)


P-CNN features [29] were created from person-centric motion and appearance features extracted
using body joint positions. The P-CNN workflow is similar to Simonyan and Zisserman’s two-stream
CNN architecture [55]. For this work, we used the publicly available P-CNN code with minor
modifications. The process of P-CNN feature extraction is shown in Figure 5. The main modules of the
process are explained below.

RGB CNN Aggregation Normalization


RGB patches Sta. Min P-CNN
Max
𝑣𝑠𝑡𝑎
Min
𝒇𝑡 Dyn. Max
𝑣𝑑𝑦𝑛 𝑣𝑎𝑝𝑝
Extract
Input
image Flow CNN Aggregation Normalization
image
patches Flow patches Sta. Min
𝑣𝑠𝑡𝑎 𝑣𝑜𝑓
Max
Min
𝒇𝑡 Dyn. Max
𝑣𝑑𝑦𝑛

Figure 5. The steps involved in a P-CNN feature descriptor calculation. For a given image patch, one
RGB and one flow CNN descriptor f t is calculated per frame t.

4.2.1. Pose Estimation


We used the latest version of OpenPose [52] for pose estimation (see Figure 4a,b). OpenPose is a
state-of-the-art method of pose estimation [56]. The pose estimator provides 18 key point estimations
(eyes, ears, nose shoulders, elbows, wrists, hips, knees, and ankles). Head position was calculated by
Drones 2019, 3, 82 10 of 16

taking the center of the eyes, ears, and nose key points. Sometimes, all of the key points in the head
were not visible (e.g., from a side view). The abdomen key point was calculated by taking the centroid
between the shoulder and hip key points. A total of 15 key points were extracted for the experiments
(head, neck, shoulders, elbows, wrists, abdomen, hips, knees, and ankles).

4.2.2. Optical Flow Calculation


Optical flow was calculated between all consecutive frames of the video using Brox et al.’s
implementation [57]. The calculated optical flow image was transformed to a heat map at the pixel
level. The velocities along x and y axes, namely, v x , vy , were mapped to the range [0, 255] by [v̂ x , v̂y ] =
s[v x , vy ] + c, with values outside the range [0, 255] truncated at 0 or 255, according to [58]. The scalar s
was calculated as s = c/a, where a = 8 was the maximum absolute value of the flow, and c = 128 was
the new center of motion velocities.

4.2.3. Extracting Part Patches


The scale of each pose was used to extract consistent sized patches from the hands, upper body,
and entire body of the subject. The scale for each pose was obtained by dividing the pose height
(in pixels) by 240. In the image coordinate system, height was calculated using the maximum difference
in y-coordinates between the head and ankles.
A region of interest around a body part was extracted by adding some margin to the top-left
and bottom-right key point positions of the body part—we called this a part patch. The margin was
calculated as C × S, where C was selected to be 100 pixels, and S was the scale. The extracted patch
was resized to 224 × 224 to match the input size of the CNN model. Figure 4 illustrates the process of
extracting part patches using an image from the Kicking class.

4.2.4. CNN Feature Aggregation


For each body part and full image, the appearance (RGB) and optical flow patches were extracted,
and their CNN features were computed using two pre-trained networks (see Figure 5). For appearance
patches, the publicly available “VGG-f” network [59] was used, whereas for optical flow patches,
the motion network from Gkioxari and Malik’s Action Tube implementation [58] was used. Both
networks have the same architecture with five convolutional and three fully-connected layers. For each
body part and frame t, a frame descriptor f t was obtained from the output of the second fully-connected
layer. Static and dynamic features were separately aggregated over time to obtain a static video
descriptor vsta and a dynamic video descriptor vdyn respectively.
Min and max aggregation schemes were used to calculate the minimum and maximum values of
each descriptor dimension k (k ∈ {1, ..., 4096}) over T frames.

mk = min f t (k),
1≤ t ≤ T
(1)
Mk = max f t (k).
1≤ t ≤ T

For static and dynamic video descriptors, the time-aggregated video descriptors were
concatenated as follows:

vsta = [m1 , . . . , mk , M1 , . . . , Mk ]> , (2)


>
vdyn = [∆m1 , . . . , ∆mk , ∆M1 , . . . , ∆Mk ] . (3)

∆ represents temporal differences in the video descriptors. The aggregated features (vsta and vdyn )
were normalized and concatenated over the number of body parts to obtain appearance features v app
and flow features vo f . The final P-CNN descriptor was obtained by concatenating v app and vo f .
Drones 2019, 3, 82 11 of 16

4.3. Performance Evaluation


The evaluation metric selected for the experiment was accuracy. Accuracy was calculated using
the scores returned by the action classifiers. We experimented with the dataset by generating three
split sets. Each split set was generated randomly with a 70 : 30 train-to-test ratio using the subject ID.
We report a mean classification accuracy of 64.36% for HLPF. As explained in Section 4.1,
classification was performed using linear kernels. Accuracies for the three split sets were 63.89%,
68.09%, and 61.11%.
The mean accuracy reported for P-CNN was 75.92%. The accuracies of the three split set were
72.22%, 81.94%, and 73.61%. Here, the classification was performed using a linear SVM.
In Table 3, we compare our dataset with two other action recognition datasets; namely, JHMDB
and MPII. Both of these datasets used Cherian et al.’s pose estimation algorithm [60]. The performance
metrics for the JHMDB and MPII datasets were taken from [29].
Two confusion matrices were calculated (see Figure 6) for assessing the action recognition
accuracies of the HLPF and P-CNN approaches. Both HLPF and P-CNN had some difficulty
distinguishing between hitting with bottle and stabbing actions, which were associated with the lowest
scores in the confusion matrix for P-CNN and noticeably low scores for HLPF. The main reason for
this was the similar patterns of body joint movements for these two classes of actions, although HLPF
does not take into account the characteristics of associated objects. Another possibility specific to
P-CNN was that the color of the bottle used in hitting with bottle was close to the background color
and sometimes difficult to identify. The classifiers can also confuse between jogging and running
actions, as the speeds of these actions varied from one actor to another. Overall, P-CNN showed better
performance for most of the classes.
As an alternative measure of accuracy, Cohen’s kappa coefficient [61] was calculated for both
the HLPF and PCNN confusion matrices. Cohen’s kappa was originally proposed in the area of
psychology as a measure of agreement between two “judges” that each classifies some objects into
p −p
some mutually exclusive categories. The coefficient was defined as κ , 1o− pee , where po is the
proportion of agreements, and pe is the proportion of chance agreements. According to Landis and
Koch [62], a κ value of < 0 should be interpreted as no agreement, 0–0.20 as slight, 0.21–0.40 as fair,
0.41–0.60 as moderate, 0.61–0.80 as substantial, and 0.81–1 as almost perfect agreement. In the context
of multi-class classification, an agreement is a true positive. Given confusion matrix [mi,j ] and k classes,

m1,1 + m2,2 + · · · + mk,k ∑kl=1 (∑kj=1 ml,j · ∑ik=1 mi,l )


po = , pe = .
∑i,j∈{1,...,k} mi,j ∑i,j∈{1,...,k} mi,j

Using Cardillo’s code [63], Cohen’s κ coefficients for the HLPF and P-CNN confusion matrices
were found to be 0.6433 and 0.7593 respectively. As per Landis and Koch’s [62] interpretation, both κ
coefficients indicated substantial agreement, with the κ for P-CNN being conspicuously higher than
that for HLPF, and in fact nearly falling in the region of almost perfect agreement.
We also performed experiments by (i) combining P-CNN and HLPF features to form a new
feature vector, and (ii) fusing the classification scores of P-CNN and HLPF classifiers. In both cases,
the performance was degraded below the accuracy reported for P-CNN alone. Therefore, we report
the P-CNN feature-based action classification as the baseline algorithm used for our dataset.

Table 3. The best reported HLPF and P-CNN action recognition results for different datasets.

Dataset Remarks HLPF P-CNN


JHMDB Res: 320 × 240, Pose [60] 25.30 61.10
MPII Cooking Res: 1624 × 1224, Pose [60] 32.60 62.30
Drone-Action Res: 1920 × 1080, OpenPose [52] 64.36 75.92
Drones 2019, 3, 82 12 of 16

1 Clapping 100 0 0 0 0 0 0 0 0 0 0 0 0 1 100 0 0 0 0 0 0 0 0 0 0 0 0


2 Hitting with bottle 0 5.6 55.6 0 0 5.6 0 0 0 22.2 0 0 11.1 2 0 11.1 33.3 0 0 0 0 0 0 55.6 0 0 0
3 Hitting with stick 0 0 77.8 0 0 0 11.1 0 0 0 0 0 11.1 3 0 0 100 0 0 0 0 0 0 0 0 0 0
4 Jogging front/back 5.6 0 0 55.6 0 0 0 33.3 0 0 5.6 0 0 4 0 0 0 66.7 0 0 0 33.3 0 0 0 0 0
5 Jogging side 0 0 0 5.6 66.7 0 0 0 5.6 0 0 22.2 0 5 0 0 0 0 66.7 0 0 0 33.3 0 0 0 0
6 Kicking 0 0 0 0 0 100 0 0 0 0 0 0 0 6 0 0 0 0 0 100 0 0 0 0 0 0 0
7 Punching 16.7 0 0 0 0 0 83.3 0 0 0 0 0 0 7 0 11.1 0 0 0 0 88.9 0 0 0 0 0 0
8 Running front/back 11.1 0 0 38.9 0 0 0 11.1 0 0 38.9 0 0 8 0 0 0 11.1 0 0 0 88.9 0 0 0 0 0
9 Running side 0 0 0 0 50 5.6 0 0 44.4 0 0 0 0 9 0 0 0 0 33.3 0 0 0 66.7 0 0 0 0
10 Stabbing 5.6 38.9 11.1 0 0 0 11.1 0 0 33.3 0 0 0 10 0 77.8 0 0 0 0 0 0 0 22.2 0 0 0
11 Walking front/back 0 0 0 0 0 0 0 0 0 0 100 0 0 11 0 0 0 0 0 0 0 0 0 0 100 0 0
12 Walking side 0 0 0 0 5.6 0 0 0 0 0 0 94.4 0 12 0 0 0 0 0 0 0 0 0 0 0 100 0
13 Waving hands 0 0 0 0 0 0 0 0 0 0 0 0 100 13 0 0 0 0 0 0 0 0 0 0 0 0 100
1 2 3 4 5 6 7 8 9 10 11 12 13 1 2 3 4 5 6 7 8 9 10 11 12 13

Figure 6. Confusion matrices calculated as percentage accuracies for HLPF (on the left) and P-CNN
(on the right).

5. Discussion
HLPF computation is based only on the spatial and temporal relationships among the key points.
It does not consider any additional information associated with the actions, such as the objects used in
the actions. In comparison, P-CNN computation includes the body parts and action-specific objects
visible in the part patches. This results in a higher action recognition accuracy for P-CNN compared to
HLPF. This observation is applicable to all three datasets used in this work.
In both HLPF and P-CNN approaches, actions with similar body joint movements can be confused
as one another, resulting in an overall reduction in accuracy. Both feature descriptors performed well
with distinctive actions. It was noted that HLPF accuracy decreased when the action was associated
with additional objects, while P-CNN was robust whether additional objects were involved or not.
We used the state-of-the-art pose estimator OpenPose to find key points. However, as with all
estimators, sometimes it missed the key points due to varying human scales and viewpoint occlusions.
The estimator can be made more robust to scale variation by increasing the input image size of the
pose estimator model, at the expense of computational overhead. Perspective distortion caused by the
use of an aerial camera is minimal in most of the videos, although in the front/back following actions,
the perspective distortion varies when the subject moves closer to or further away from the drone.
This distortion can be compensated for to some extent using an approach similar to that reported by
Perera et al. [64].
This dataset contains only single-person actions. As explained in Sections 1 and 2, drone-based
human action recognition is still an emerging research area. The closest research area to this is
gesture-based drone control. We believe that single-person actions provide an opportunity for many
researchers to start or advance their research in aerial action recognition. That is why this dataset
focuses on the first step in aerial action recognition: single-person action recognition. Our ongoing
research includes extending this dataset to a multi-person version.
In multi-person action recognition studies, normally, the actions are broadly categorized based
on the event. A relevant dataset and approach are described in the VIRAT aerial dataset [13]. When
detecting the actions of a group of people, each individual in the scene should be detected and analyzed
for the collective or individual actions [16]. However, depending on the scene, multiple subjects can
improve the detection accuracy in the actions (e.g., fighting).
This dataset has been created to help progress research in aerial action recognition. The work
presented in this paper is intended to provide a sufficient amount of data to cover some common
human actions. These actions can be visible from a low-altitude slow-flying drone. When recording
videos, we used a relatively clutter-free background and maximized the image details by keeping the
subject roughly in the middle of the frame throughout the drone’s hover and follow modes. Therefore,
our dataset does not include videos with complex backgrounds and challenging flight maneuvers.

6. Conclusions
We presented an action recognition dataset, called Drone-Action, recorded by a hovering drone.
The dataset contains 240 HD videos lasting for a total of 44.6 minutes. The dataset was prepared
Drones 2019, 3, 82 13 of 16

using 13 selected actions from common outdoor human actions. The actions were recorded from 10
participants in an outdoor setting while the drone was in hover or follow mode. The rich variation of
body size, camera motion, and phase, makes our dataset challenging for action recognition. The dataset
contains videos recorded from a low-altitude, slow-flying drone. The action classes, the relatively large
images of human subjects and high resolution image details extend the dataset’s applicability to a
wider research community.
We evaluated this new dataset using HLPF and P-CNN descriptors, and reported an overall
baseline action recognition accuracy of 75.92% using P-CNN. This dataset is useful for research
involving surveillance, situational awareness, general action recognition, violence detection, and gait
recognition. The Drone-Action dataset is available at https://asankagp.github.io/droneaction.

Author Contributions: The original draft preparation was done by A.G.P. Both Y.W.L.’s and J.C. contributions
were to aid in conceptualization, reviewing, and editing.
Funding: This project was partly supported by Project Tyche, the Trusted Autonomy Initiative of the Defence
Science and Technology Group (grant number myIP6780).
Acknowledgments: We thank the student volunteers who participated in the data collection.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Gonçalves, J.; Henriques, R. UAV photogrammetry for topographic monitoring of coastal areas. ISPRS J.
Photogramm. Remote Sens. 2015, 104, 101–111. [CrossRef]
2. Barbedo, J.G.A. A Review on the Use of Unmanned Aerial Vehicles and Imaging Sensors for Monitoring
and Assessing Plant Stresses. Drones 2019, 3, 40. [CrossRef]
3. Al-Kaff, A.; Moreno, F.M.; José, L.J.S.; García, F.; Martín, D.; de la Escalera, A.; Nieva, A.; Garcéa, J.L.M.
VBII-UAV: Vision-Based Infrastructure Inspection-UAV. In Recent Advances in Information Systems and
Technologies; Rocha, Á., Correia, A.M., Adeli, H., Reis, L.P., Costanzo, S., Eds.; Springer International
Publishing: Cham, Switzerland, 2017; pp. 221–231.
4. Erdelj, M.; Natalizio, E.; Chowdhury, K.R.; Akyildiz, I.F. Help from the Sky: Leveraging UAVs for Disaster
Management. IEEE Pervasive Comput. 2017, 16, 24–32. [CrossRef]
5. Peschel, J.M.; Murphy, R.R. On the Human–Machine Interaction of Unmanned Aerial System Mission
Specialists. IEEE Trans. Hum.-Mach. Syst. 2013, 43, 53–62. [CrossRef]
6. Chahl, J. Unmanned Aerial Systems (UAS) Research Opportunities. Aerospace 2015, 2, 189–202. [CrossRef]
7. Xia, G.S.; Bai, X.; Ding, J.; Zhu, Z.; Belongie, S.; Luo, J.; Datcu, M.; Pelillo, M.; Zhang, L. DOTA: A Large-Scale
Dataset for Object Detection in Aerial Images. In Proceedings of the IEEE Conference on Computer Vision
and Pattern Recognition (CVPR), Salt Lake City, UT, USA, 18–22 June 2018.
8. Razakarivony, S.; Jurie, F. Vehicle detection in aerial imagery : A small target detection benchmark. J. Vis.
Commun. Image Represent. 2016, 34, 187–203. [CrossRef]
9. Krajewski, R.; Bock, J.; Kloeker, L.; Eckstein, L. The highD Dataset: A Drone Dataset of Naturalistic Vehicle
Trajectories on German Highways for Validation of Highly Automated Driving Systems. In Proceedings
of the 2018 21st International Conference on Intelligent Transportation Systems (ITSC), Maui, HI, USA,
4–7 November 2018; pp. 2118–2125. [CrossRef]
10. Interstate 80 Freeway Dataset. 2019. Available online: :https://www.fhwa.dot.gov/publications/research/
operations/06137/index.cfm (accessed on 2 November 2019).
11. Zhu, P.; Wen, L.; Bian, X.; Haibin, L.; Hu, Q. Vision Meets Drones: A Challenge. arXiv 2018, arXiv:1804.07437.
12. Carletti, V.; Greco, A.; Saggese, A.; Vento, M. Multi-Object Tracking by Flying Cameras Based on a
Forward-Backward Interaction. IEEE Access 2018, 6, 43905–43919. [CrossRef]
13. Oh, S.; Hoogs, A.; Perera, A.; Cuntoor, N.; Chen, C.C.; Lee, J.T.; Mukherjee, S.; Aggarwal, J.K.; Lee, H.;
Davis, L.; et al. A large-scale benchmark dataset for event recognition in surveillance video. In Proceedings
of the CVPR 2011, Providence, RI, USA, 20–25 June 2011; pp. 3153–3160. [CrossRef]
14. University of Central Florida. UCF-ARG Data Set. 2011. Available online: http://crcv.ucf.edu/data/UCF-
ARG.php (accessed on 2 November 2019).
Drones 2019, 3, 82 14 of 16

15. Mueller, M.; Smith, N.; Ghanem, B. A Benchmark and Simulator for UAV Tracking. In Proceedings of
the Computer Vision—ECCV 2016, Amsterdam, The Netherlands, 8–16 October 2016; Leibe, B., Matas, J.,
Sebe, N., Welling, M., Eds.; Springer International Publishing: Cham, Switzerland, 2016; pp. 445–461.
16. Barekatain, M.; Martí, M.; Shih, H.F.; Murray, S.; Nakayama, K.; Matsuo, Y.; Prendinger, H. Okutama-Action:
An Aerial View Video Dataset for Concurrent Human Action Detection. In Proceedings of the 2017
IEEE Conference on Computer Vision and Pattern Recognition Workshops (CVPRW), Honolulu, HI, USA,
21–26 July 2017; pp. 2153–2160. [CrossRef]
17. Perera, A.G.; Wei Law, Y.; Chahl, J. UAV-GESTURE: A Dataset for UAV Control and Gesture Recognition.
In Proceedings of the European Conference on Computer Vision (ECCV) Workshops, Munich, Germany,
8–14 September 2018; pp. 117–128. [CrossRef]
18. Natarajan, K.; Nguyen, T.D.; Mete, M. Hand Gesture Controlled Drones: An Open Source Library.
In Proceedings of the 2018 1st International Conference on Data Intelligence and Security (ICDIS), South
Padre Island, TX, USA, 8–10 April 2018; pp. 168–175. [CrossRef]
19. Lee, J.; Tan, H.; Crandall, D.; Šabanović, S. Forecasting Hand Gestures for Human-Drone Interaction.
In Proceedings of the Companion of the 2018 ACM/IEEE International Conference on Human-Robot
Interaction, Chicago, IL, USA, 5–8 March 2018; pp. 167–168. [CrossRef]
20. Hsu, H.J.; Chen, K.T. DroneFace: An Open Dataset for Drone Research. In Proceedings of the 8th ACM on
Multimedia Systems Conference, Taipei, Taiwan, 20–23 June 2017; pp. 187–192. [CrossRef]
21. Kalra, I.; Singh, M.; Nagpal, S.; Singh, R.; Vatsa, M.; Sujit, P.B. DroneSURF: Benchmark Dataset for
Drone-based Face Recognition. In Proceedings of the 2019 14th IEEE International Conference on Automatic
Face Gesture Recognition (FG 2019), Lille, France, 14–18 May 2019; pp. 1–7. [CrossRef]
22. Carletti, V.; Greco, A.; Saggese, A.; Vento, M. An intelligent flying system for automatic detection of faults in
photovoltaic plants. J. Ambient Intell. Hum. Comput. 2019. [CrossRef]
23. Avola, D.; Cinque, L.; Foresti, G.L.; Martinel, N.; Pannone, D.; Piciarelli, C. A UAV Video Dataset for
Mosaicking and Change Detection From Low-Altitude Flights. IEEE Trans. Syst. Man Cybern. Syst. 2018.
[CrossRef]
24. Sensefly Mosaic Datasets. 2019. Available online: https://www.sensefiy.com/drones/example-datasets.
html (accessed on 2 November 2019).
25. Lottes, P.; Khanna, R.; Pfeifer, J.; Siegwart, R.; Stachniss, C. UAV-based crop and weed classification for smart
farming. In Proceedings of the 2017 IEEE International Conference on Robotics and Automation (ICRA),
Singapore, 29 May–3 June 2017; pp. 3024–3031. [CrossRef]
26. Monteiro, A.; von Wangenheim, A. Orthomosaic Dataset of RGB Aerial Images for Weed Mapping. 2019.
Available online: http://www.lapix.ufsc.br/weed-mapping-sugar-cane (accessed on 2 November 2019).
27. Herath, S.; Harandi, M.; Porikli, F. Going deeper into action recognition: A survey. Image Vis. Comput. 2017,
60, 4–21. [CrossRef]
28. Mabrouk, A.B.; Zagrouba, E. Abnormal behavior recognition for intelligent video surveillance systems: A
review. Expert Syst. Appl. 2018, 91, 480–491. [CrossRef]
29. Cheron, G.; Laptev, I.; Schmid, C. P-CNN: Pose-Based CNN Features for Action Recognition. In Proceedings
of the IEEE International Conference on Computer Vision (ICCV), Santiago, Chile, 7–13 December 2015.
30. Jhuang, H.; Gall, J.; Zuffi, S.; Schmid, C.; Black, M.J. Towards Understanding Action Recognition.
In Proceedings of the 2013 IEEE International Conference on Computer Vision, Sydney, Australia,
1–8 December 2013; pp. 3192–3199. [CrossRef]
31. Schuldt, C.; Laptev, I.; Caputo, B. Recognizing human actions: A local SVM approach. In Proceedings of the
17th International Conference on Pattern Recognition, Cambridge, UK, 26 August 2004; Volume 3, pp. 32–36.
32. Blank, M.; Gorelick, L.; Shechtman, E.; Irani, M.; Basri, R. Actions as space-time shapes. In Proceedings of
the Tenth IEEE International Conference on Computer Vision (ICCV’05), Beijing, China, 17–21 October 2005;
Volume 2, pp. 1395–1402. [CrossRef]
33. Soomro, K.; Zamir, A.R.; Shah, M. UCF101: A Dataset of 101 Human Actions Classes From Videos in The Wild;
Technical Report; UCF Center for Research in Computer Vision: Orlando, FL, USA, 2012.
34. Zhang, W.; Zhu, M.; Derpanis, K.G. From Actemes to Action: A Strongly-Supervised Representation for
Detailed Action Understanding. In Proceedings of the 2013 IEEE International Conference on Computer
Vision, Sydney, Australia, 1–8 December 2013; pp. 2248–2255. [CrossRef]
Drones 2019, 3, 82 15 of 16

35. Karpathy, A.; Toderici, G.; Shetty, S.; Leung, T.; Sukthankar, R.; Fei-Fei, L. Large-Scale Video Classification
with Convolutional Neural Networks. In Proceedings of the 2014 IEEE Conference on Computer Vision and
Pattern Recognition, Columbus, OH, USA, 24–27 June 2014; pp. 1725–1732. [CrossRef]
36. Heilbron, F.C.; Escorcia, V.; Ghanem, B.; Niebles, J.C. ActivityNet: A large-scale video benchmark for human
activity understanding. In Proceedings of the 2015 IEEE Conference on Computer Vision and Pattern
Recognition (CVPR), Boston, MA, USA, 7–12 June 2015; pp. 961–970. [CrossRef]
37. Feichtenhofer, C.; Pinz, A.; Zisserman, A. Convolutional Two-Stream Network Fusion for Video Action
Recognition. In Proceedings of the 2016 IEEE Conference on Computer Vision and Pattern Recognition
(CVPR), Las Vegas, NV, USA, 27–30 June 2016; pp. 1933–1941. [CrossRef]
38. Abu-El-Haija, S.; Kothari, N.; Lee, J.; Natsev, A.; Toderici, G.; Varadarajan, B.; Vijayanarasimhan, S.
YouTube-8M: A Large-Scale Video Classification Benchmark. arXiv 2016, arXiv:1609.08675.
39. Kay, W.; Carreira, J.; Simonyan, K.; Zhang, B.; Hillier, C.; Vijayanarasimhan, S.; Viola, F.; Green, T.; Back, T.;
Natsev, A.; et al. The Kinetics Human Action Video Dataset. arXiv 2017, arXiv:1705.06950.
40. Zhao, H.; Yan, Z.; Torresani, L.; Torralba, A. HACS: Human Action Clips and Segments Dataset for
Recognition and Temporal Localization. arXiv 2019, arXiv:1712.09374.
41. Du, D.; Qi, Y.; Yu, H.; Yang, Y.; Duan, K.; Li, G.; Zhang, W.; Huang, Q.; Tian, Q. The Unmanned Aerial Vehicle
Benchmark: Object Detection and Tracking. In Proceedings of the European Conference on Computer Vision
(ECCV), Munich, Germany, 8–14 September 2018.
42. Song, Y.; Demirdjian, D.; Davis, R. Tracking body and hands for gesture recognition: NATOPS aircraft
handling signals database. Face Gesture 2011, 500–506. [CrossRef]
43. University of Central Florida. UCF Aerial Action Dataset. 2011. Available online: http://crcv.ucf.edu/data/
UCF_Aerial_Action.php (accessed on 02 Novomber 2019).
44. Bonetto, M.; Korshunov, P.; Ramponi, G.; Ebrahimi, T. Privacy in mini-drone based video surveillance.
In Proceedings of the 2015 11th IEEE International Conference and Workshops on Automatic Face and
Gesture Recognition (FG), Ljubljana, Slovenia, 4–8 May 2015; Volume 4, pp. 1–6. [CrossRef]
45. Ovtcharov, K.; Ruwase, O.; Kim, J.Y.; Fowers, J.; Strauss, K.; Chung, E.S. Accelerating deep convolutional
neural networks using specialized hardware. Microsoft Res. Whitepaper 2015, 2, 1–4.
46. Rudol, P.; Doherty, P. Human Body Detection and Geolocalization for UAV Search and Rescue Missions
Using Color and Thermal Imagery. In Proceedings of the 2008 IEEE Aerospace Conference, Big Sky, MT,
USA, 1–8 March 2008; pp. 1–8. [CrossRef]
47. Oreifej, O.; Mehran, R.; Shah, M. Human identity recognition in aerial images. In Proceedings of the
2010 IEEE Conference on Computer Vision and Pattern Recognition (CVPR), San Francisco, CA, USA,
13–18 June 2010; pp. 709–716. [CrossRef]
48. Yeh, M.C.; Chiu, H.K.; Wang, J.S. Fast medium-scale multiperson identification in aerial videos.
Multimed. Tools Appl. 2016, 75, 16117–16133. [CrossRef]
49. Al-Naji, A.; Perera, A.G.; Chahl, J. Remote monitoring of cardiorespiratory signals from a hovering
unmanned aerial vehicle. BioMedical Eng. OnLine 2017, 16, 101. [CrossRef]
50. De Souza, F.D.; Chavez, G.C.; do Valle, E.A., Jr.; Araújo, A.D.A. Violence Detection in Video Using
Spatio-Temporal Features. In Proceedings of the 2010 23rd SIBGRAPI Conference on Graphics, Patterns and
Images, Gramado, Brazil, 30 August–3 September 2010; pp. 224–230. [CrossRef]
51. Datta, A.; Shah, M.; Lobo, N.D.V. Person-on-person violence detection in video data. In Proceedings
of the Object Recognition Supported by User Interaction for Service Robots, Quebec City, QC, Canada,
11–15 August 2002. [CrossRef]
52. Cao, Z.; Simon, T.; Wei, S.E.; Sheikh, Y. Realtime Multi-Person 2D Pose Estimation using Part Affinity Fields.
In Proceedings of the CVPR, Honolulu, HI, USA, 21–26 July 2017.
53. Ryoo, M.S.; Aggarwal, J.K. Spatio-temporal relationship match: Video structure comparison for recognition
of complex human activities. In Proceedings of the 2009 IEEE 12th International Conference on Computer
Vision, Kyoto, Japan, 29 September–2 October 2009; pp. 1593–1600. [CrossRef]
54. Robicquet, A.; Sadeghian, A.; Alahi, A.; Savarese, S. Learning Social Etiquette: Human Trajectory
Understanding In Crowded Scenes. In Proceedings of the Computer Vision—ECCV, Amsterdam, The
Netherlands, 11–14 October 2016; Leibe, B., Matas, J., Sebe, N., Welling, M., Eds.; Springer International
Publishing: Cham, Switzerland, 2016; pp. 549–565.
Drones 2019, 3, 82 16 of 16

55. Simonyan, K.; Zisserman, A. Two-Stream Convolutional Networks for Action Recognition in Videos.
In Proceedings of the 27th International Conference on Neural Information Processing Systems, Montreal,
QC, Canada, 8–13 December 2014; pp. 568–576.
56. Zhao, M.; Li, T.; Alsheikh, M.A.; Tian, Y.; Zhao, H.; Torralba, A.; Katabi, D. Through-Wall Human Pose
Estimation Using Radio Signals. In Proceedings of the 2018 IEEE/CVF Conference on Computer Vision and
Pattern Recognition, Salt Lake City, UT, USA, 18-22 June 2018; pp. 7356–7365. [CrossRef]
57. Brox, T.; Bruhn, A.; Papenberg, N.; Weickert, J. High Accuracy Optical Flow Estimation Based on a
Theory for Warping. In Proceedings of the Computer Vision—ECCV 2004; Pajdla, T., Matas, J., Eds.; Springer:
Berlin/Heidelberg, Germany, 2004; pp. 25–36.
58. Gkioxari, G.; Malik, J. Finding Action Tubes. In Proceedings of the IEEE Conference on Computer Vision
and Pattern Recognition (CVPR), Boston, MA, USA, 7–12 June 2015.
59. Chatfield, K.; Simonyan, K.; Vedaldi, A.; Zisserman, A. Return of the Devil in the Details: Delving Deep into
Convolutional Nets. arXiv 2014, arXiv:1405.3531v4.
60. Cherian, A.; Mairal, J.; Alahari, K.; Schmid, C. Mixing Body-Part Sequences for Human Pose
Estimation. In Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition (CVPR),
Columbus, OH, USA, 24–27 June 2014.
61. Cohen, J. A coefficient of agreement for nominal scales. Educ. Psychol. Meas. 1960, 20, 37–46. [CrossRef]
62. Landis, J.R.; Koch, G.G. The measurement of observer agreement for categorical data. Biometrics 1977, 33,
159–174. [CrossRef] [PubMed]
63. Cardillo, G. Compute the Cohen’s Kappa (Version 2.0.0.0). 2018. Available online: http://www.mathworks.
com/matlabcentral/fileexchange/15365 (accessed on 2 November 2019).
64. Perera, A.G.; Law, Y.W.; Chahl, J. Human Pose and Path Estimation from Aerial Video Using Dynamic
Classifier Selection. Cogn. Comput. 2018, 10, 1019–1041. [CrossRef]

c 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like